twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the...

19
INVITED REVIEW Twenty odd years of stretch-sensitive channels O. P. Hamill Received: 14 June 2006 / Accepted: 27 June 2006 / Published online: 21 September 2006 # Springer-Verlag 2006 Abstract After formation of the giga-seal, the membrane patch can be stimulated by hydrostatic or osmotic pressure gradients applied across the patch. This feature led to the discovery of stretch-sensitive or mechanosensitive (MS) channels, which are now known to be ubiquitously expressed in cells representative of all the living kingdoms. In addition to mechanosensation, MS channels have been implicated in many basic cell functions, including regula- tion of cell volume, shape, and motility. The successful cloning, overexpression, and crystallization of bacterial MS channel proteins combined with patch clamp and modeling studies have provided atomic insight into the working of these nanomachines. In particular, studies of MS channels have revealed new understanding of how the lipid bilayer modulates membrane protein function. Three major mem- brane protein families, transient receptor potential, 2 pore domain K + , and the epithelial Na + channels, have been shown to form MS channels in animal cells, and their polymodal activation embrace fields far beyond mechano- sensitivity. The discovery of new drugs highly selective for MS channels (mechanopharmaceutics) and the demon- stration of MS channel involvement in several major human diseases (mechanochannelopathies) provide added moti- vation for devising new techniques and approaches for studying MS channels. Keywords Mechanosensitive channels . Mechanotransduction . Transient receptor potential . Patch clamp . Giga seal Introduction The formation of the giga-seal depends upon applying suction to draw the membrane into the pipette. Once the seal is formed, suction is not necessary, and it should be released [53]. However, the membrane patch spanning the pipette can now be mechanically stimulated by hydrostatic or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of cell-swelling and stretch-activated channel currents [50, 52]. Over the last 20-odd years, interest in mechanosensitive (MS) channels has progressed from being a possible patch clamp recording artifact to a central player in our understanding of proteinbilayer interactions and a promising new therapeutic target against several major human diseases. This article highlights some recent devel- opments and unresolved issues regarding MS channels, with a major focus on the MS Ca 2+ permeant cation channel (MscCa) recently identified in vertebrate cells [97]. What happens to the membrane patch in the pipette? An important issue for MS channels is how the process of aspiration and sealing of the membrane in the pipette alters the mechanics and possible stretch sensitivity of channels in the patch. Because of the small size and inaccessibility of the patch in the pipette, a variety of techniques, including high-resolution video imaging [121, 152154, 184], high-voltage electron microscopy [142], atomic force microscopy [70], and fluorescence-imaged microdeformation [32, 33] have been used to study the aspirated patch and its underlying cell cytoskeleton (CSK). Here we focus on results obtained on the Xenopus oocyte [182185]. The first issue is whether the pressure/ suction applied to the patch after seal formation somehow Pflugers Arch - Eur J Physiol (2006) 453:333351 DOI 10.1007/s00424-006-0131-0 O. P. Hamill (*) Neurosciences and Cell Biology, UTMB, Galveston, TX 77555, USA e-mail: [email protected]

Upload: others

Post on 19-Jun-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

INVITED REVIEW

Twenty odd years of stretch-sensitive channels

O. P. Hamill

Received: 14 June 2006 /Accepted: 27 June 2006 / Published online: 21 September 2006# Springer-Verlag 2006

Abstract After formation of the giga-seal, the membranepatch can be stimulated by hydrostatic or osmotic pressuregradients applied across the patch. This feature led to thediscovery of stretch-sensitive or mechanosensitive (MS)channels, which are now known to be ubiquitouslyexpressed in cells representative of all the living kingdoms.In addition to mechanosensation, MS channels have beenimplicated in many basic cell functions, including regula-tion of cell volume, shape, and motility. The successfulcloning, overexpression, and crystallization of bacterial MSchannel proteins combined with patch clamp and modelingstudies have provided atomic insight into the working ofthese nanomachines. In particular, studies of MS channelshave revealed new understanding of how the lipid bilayermodulates membrane protein function. Three major mem-brane protein families, transient receptor potential, 2 poredomain K+, and the epithelial Na+ channels, have beenshown to form MS channels in animal cells, and theirpolymodal activation embrace fields far beyond mechano-sensitivity. The discovery of new drugs highly selective forMS channels (“mechanopharmaceutics”) and the demon-stration of MS channel involvement in several major humandiseases (“mechanochannelopathies”) provide added moti-vation for devising new techniques and approaches forstudying MS channels.

Keywords Mechanosensitive channels .

Mechanotransduction . Transient receptor potential .

Patch clamp . Giga seal

Introduction

The formation of the giga-seal depends upon applyingsuction to draw the membrane into the pipette. Once theseal is formed, suction is not necessary, and it should bereleased [53]. However, the membrane patch spanning thepipette can now be mechanically stimulated by hydrostaticor osmotic pressure gradients applied across the patchwithout disrupting the seal. This feature allowed the firstrecordings of cell-swelling and stretch-activated channelcurrents [50, 52]. Over the last 20-odd years, interest inmechanosensitive (MS) channels has progressed from beinga possible patch clamp recording artifact to a central playerin our understanding of protein–bilayer interactions and apromising new therapeutic target against several majorhuman diseases. This article highlights some recent devel-opments and unresolved issues regarding MS channels,with a major focus on the MS Ca2+ permeant cationchannel (MscCa) recently identified in vertebrate cells [97].

What happens to the membrane patch in the pipette?

An important issue for MS channels is how the process ofaspiration and sealing of the membrane in the pipettealters the mechanics and possible stretch sensitivity ofchannels in the patch. Because of the small size andinaccessibility of the patch in the pipette, a variety oftechniques, including high-resolution video imaging [121,152–154, 184], high-voltage electron microscopy [142],atomic force microscopy [70], and fluorescence-imagedmicrodeformation [32, 33] have been used to study theaspirated patch and its underlying cell cytoskeleton(CSK). Here we focus on results obtained on the Xenopusoocyte [182–185]. The first issue is whether the pressure/suction applied to the patch after seal formation somehow

Pflugers Arch - Eur J Physiol (2006) 453:333–351DOI 10.1007/s00424-006-0131-0

O. P. Hamill (*)Neurosciences and Cell Biology, UTMB,Galveston, TX 77555, USAe-mail: [email protected]

Page 2: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

induces changes in the seal resistance that appear as “MS-channel-like” events. This idea is not entirely far-fetchedbecause video imaging indicates that suction tends to peal themembrane off the walls of the pipette [121], and gated,cation-selective channels have been recorded with patchpipettes sealed onto noncellular hydrophobic surfaces [145].However, although strong suction can rupture the patch, ittypically does not disrupt the giga-seal, thereby allowing fortight seal whole-cell and/or outside-out patch recording [53].In addition, suction ramps applied to cell-attached frogoocyte patches reversibly activate either a saturating macro-scopic current (Fig. 1a) or a unitary amplitude current event(Fig. 1b) depending upon patch area [60]. These currentwaveforms are consistent with multiple and single MSchannel patches, respectively, but difficult to reconcile withMS changes in seal resistance. Even more compelling is thatpatches formed on pure liposomes fail to express MSchannel currents [97, 102, 122, 160]. This absence allowedfor the identification of MS channel proteins followingfunctional reconstitution of solubilized membrane proteinsfrom bacteria and archaea [81, 158, 159] and, most recently,from Xenopus oocytes [97].

Although MS channels are clearly not seal “leaks,” thesealing process does change patch geometry and theunderlying CSK, thereby altering patch mechanics. Forexample, Fig. 2a and b shows electron microscopy (EM)

images of the Xenopus oocyte surface, indicating itsextensive membrane folding and high density of microvilli(est. ∼7 microvilli per square micrometer) [181, 184, 185].This complex membrane geometry is also reflected inelectrical capacitance (Cm) measurements that indicate amembrane surface area that is tenfold greater than thatrequired for the cell’s volume [184]. In contrast, high-resolution video images of cell-attached oocyte patches(Fig. 2c, d) indicate an optically smooth membrane that ispulled flat and perpendicular to the walls of the pipette[184; see also 113, 152, 153, 161]. Furthermore, Cm

measurements indicate a patch area of approximately50 μm2, consistent with the patch geometry but inconsistentwith the approximately 500 μm2 expected if the approxi-mately 300 microvilli and membrane folds evident inFig. 2a and b were preserved during the sealing process[184]. Presumably, the suction used to obtain the giga-sealis sufficient to smooth out the surface folds and microvilliso that the cell membrane is now tightly stretched over anexpanded CSK (Fig. 3). This smoothing out of microvilli isnot an exclusive patch clamp phenomenon because asimilar phenomenon has been visualized in EM studies ofcells undergoing osmotic inflation [82] and spreadingbefore cell migration [37]. In these cases, the process ispresumably reversible, indicating the considerable plasticityof the microvilli and their supporting CSK.

There are at least two related mechanisms by whichchanges in patch geometry will increase stretch sensitivityof channels in the patch. First, in the absence of any excessmembrane, brief pressure pulses applied to the patch willrapidly flex the membrane and increase bilayer tension(Fig. 2c, d). The flexing of the membrane either outwardwith suction or inward with pressure results in the rapidactivation (<1 ms) of inward channel currents [107, 108,184]. In contrast, more sustained pressures would berequired to inflate the oocyte and smooth out the membranereserves to increase membrane tension (Tm) [15, 117, 184].Second, according to Laplace’s law, P=2 Tm/r, the pressure(P) required to activate channels in the patch with a radiusof curvature (r) of approximately 1 μm would be 1/20th ofthat required to activate the same channels located onmicrovilli that have a radii of curvature of approximately0.05 μm. For example, stimulus–peak current relationsshown in Fig. 4 indicate that half the channels in an oocytepatch are activated by a suction of approximately 10 mmHg(∼1.3 kN m−2), which translates to a tension of approxi-mately 0.6 mN m−1 (i.e., the near-symmetrical suction/pressure relations indicate a tension-gated channel andjustify Laplace’s law). To achieve the same tension inmicrovilli would require a suction of 200 mmHg. However,this would exceed the approximately 100-mmHg thatcauses patch rupture (i.e., a lytic tension of ∼6 mN m−1)under these conditions [119].

Fig. 1 Ramps of suction applied to two different-size patches formedon a Xenopus oocyte are consistent with a finite number of discreteMS channels but inconsistent with pressure-induced leaks in the seal.a A relatively large patch formed with an approximately 3-μm-diameter tip pipette shows a current that fully saturates at around40 mmHg. The arrow indicates the initial opening of 5 pA singleMscCa. b A smaller patch formed with an approximately 0.50-μm-diameter tip pipette reveals the opening of single MscCa, indicatingthat once open, the channel current was independent of suction. Theseresults indicate that MscCa is either open or closed, and the saturationof current in a multichannel patch represents the balance betweenopening and closing rates for a finite number of channels

334 Pflugers Arch - Eur J Physiol (2006) 453:333–351

Page 3: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

The sealing process may also alter patch mechanics bychanging the CSK structure underlying the patch [55, 63,173]. When a gentle sealing protocol is used to achievethe giga-seal, the cortical CSK network that is pulled into thepipette may be dilated without disrupting its links withthe membrane [54, 185; see also 32, 33]. However, ifthe suction used to draw the membrane into the pipetteexceeds the strength of CSK–membrane linkages, then aCSK-free membrane (bleb) may be formed [55, 63, 112].Similarly, after the membrane has sealed in the pipette,additional suction may cause the blebbing at the membranecap as shown in Fig. 5 [185]. This membrane blebbing caneither increase or decrease the stretch sensitivity depending

upon the specific MS channel. For example, the TREK andTRAAK MscK channels show an increase in stretchsensitivity, presumably because the CSK normally acts as aconstraint and prevents tension being conveyed to the bilayer[69, 151]. On the other hand, MscCa typically shows a lossof both stretch sensitivity and fast dynamics presumablybecause they depend upon CSK interactions with thechannel/membrane (Fig. 5) [55, 156, 185].

In summary, giga-seal formation introduces significantchanges in patch mechanics that can alter the mechano-sensitivity of channels in the patch. The extrinsic changesin membrane geometry and CSK structure may havedifferent effects on specific channels depending upon

Fig. 2 Comparison of themembrane geometry of theoocyte surface and of the patchsealed in a pipette. a Transmis-sion EM of the oocyte surfaceshowing prominent microvillicontaining dark cytoplasmicmaterial. b Scanning EM of theoocyte surface, indicating thehigh density of microvilli.c High-resolution video imagesof a membrane patch before(0 ms), during (50 ms), and after(250 ms) a 100-ms suction step.d The same patch as in C excepta 100-ms pressure step wasapplied. Both suction and pres-sure steps activated a 50-pAinward current (modified fromZhang and Hamill [184] andZhang et al. [185] withpermission)

Pflugers Arch - Eur J Physiol (2006) 453:333–351 335

Page 4: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

their intrinsic properties (i.e., protein structure and protein–CSK interactions).

Do animal cells generate/experience membrane tensionsthat activate MS channels?

Although it has been questioned whether animal cellsexperience the same in-plane membrane tensions thatactivate MS channels in the patch [30, 149, 172], there isevidence that they can experience even larger tensions thatresult in membrane rupture. In particular, cell woundingevents, as judged by the cell filling with the membraneimpermeant fluorescein-labeled dextran, have been ob-served in experimentally unperturbed rodent skin, gutepithelium, cardiac, and skeletal muscle, and found toincrease in frequency with mechanical loading [11, 109–111, 164]. The cell types that commonly experiencewounding in vivo include epidermal cells and fibroblasts(skin), epithelial cells and smooth muscle (GI tract andrespiratory system) endothelial cells, and smooth muscle(cardiovascular system) and peripheral neurons. The pro-portion of the cells wounded in the various systems canrange from 2 to as high as 25%. In eccentrically exercisedmuscle (e.g., downhill running), there can be a tenfold

increase in cell wounding compared with resting muscle,and in dystrophic mice that lack the CSK–structuralprotein dystrophin, exercise can produce massive wound-ing events that ultimately overload the muscle regene-ration mechanisms [180]. Normal migrating cells cangenerate traction forces that not only lengthen and stretchthe cell but also cause cell fragments to be ripped off anddeposited along the migration trail [104]. Furthermore, ifthe normal contractile mechanisms that allow a migratingcell to retract its rear are blocked, the front of the cell cantear away from the cell body and move off as a motile cellfragment [170]. The common occurrence of these trau-matic mechanical stresses under physiological and path-ological conditions has presumably provided strongselective pressure for the evolution of the membraneresealing mechanism(s) that is widely expressed ineukaryotic cells [11, 109–111, 164].

Despite the above evidence of membrane rupturetensions (i.e., >5 mN m−1), direct estimates of membranetension in “resting” isolated cells indicate much lowervalues (<0.1 mN m−1) [28–31, 116, 138]. The tensionmeasurement involves pulling a tether from the cell surfaceusing optical tweezers and measuring the force required tomaintain it at constant length. The basic assumption is thatmembrane tension is contiguous over the whole surface sothat pulling a tether from one region perturbs the tension inall regions of the cell membrane. From the static tetherforce (F0) measured in piconewtons, one can estimate themembrane tension Tm according to the equation:

F0 ¼ 2π 2BTmð Þ1 2=

where the membrane bending stiffness (B) is assumed to beconstant with a value of 2.7×10−19 N m−1. The practicallimitation of this technique is that the optical tweezers canonly sustain forces up to 100 pN, which would correspondto a tension of 0.5 mN m−1. For an animal cell with itscortical CSK, the measured tension is assumed to representa combination of in-plane tension and CSK adhesion, and isreferred to as the “apparent” tension. However, Tmmeasurements of membrane blebs that lack CSK indicatethe in-plane tension contributes only 25% of the Tm value[28]. Experiments on two different cell types (RBL 2H3cells and snail neurons) indicate that cell swelling increasessteady-state tensions from approximately 0.04 to 0.12mN m−1,which then returns to approximately 0.04mN m−1 with reshrinking [30, 31]. However, the samecells also experience tension surges that exceed the strengthof the trap (i.e., >0.5 mN m−1). Based on other experimentsmeasuring exocytosis/endocytosis as a function of mem-brane tension, it has been proposed that membrane surfacearea and tension are in a feedback loop in which hightensions favor membrane recruitment, and low tensionfavors membrane retrieval [31, 116, 148]. As a conse-

Fig. 3 Schematic illustrating the proposed smoothing out of micro-villi caused by the pipette aspiration and giga-seal formation

336 Pflugers Arch - Eur J Physiol (2006) 453:333–351

Page 5: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

quence, it has been proposed that surface area regulation(SAR) maintains membrane tension around a relatively lowset point of approximately 0.1 mN m−1, which would bewell below the lytic tension (≥5 mN m−1) and the near-lytictensions (∼4 mN m−1) required to activate the bacterial MSchannels. However, the direct measurement of high tensionsurges exceeding the low set point and the occurrence ofcell wounding events indicate that SAR mechanisms can besaturated. Furthermore, as indicated in Fig. 6, lowertensions are required to activate MS channels in animalscells (T50% for MscK=2.4 mN m−1) and (MscCa ∼0.6mN m−1) compared with MscL (∼4.7 mN m−1) thatfunctions as a last-resort safety valve [94]. As a conse-

quence, MS channels in animal cells would seem moregeared to regulating processes with lower tension set pointssuch as regulatory volume decrease [20, 26, 146, 147, 169]cell locomotion [92] and, possibly, SAR via Ca2+-inducedexocytosis. However, it appears that integrins rather thanMscCa act as the mechanosensor for MS exocytosis/membrane trafficking at the frog neuromuscular junction[24] and the oocyte [96].

“Mechanopharmaceutics”

Progress in the MS channel field would be greatlyenhanced by the discovery of high-affinity agents that

Fig. 4 Inward current responsesto suction and pressure pulsesapplied to an oocyte patch.a Both suction and pressurepulses (2.5 s) result in rapidopening of MscCa that mostlyclose within 200 ms of thepulse. b Schematic showingflexion of the patch outward(suction) or inward (pressure).c Stimulus–response relationsfor suction and pressure steps.The sigmoid fits indicate thatsuction (P0.5=−10 mmHg) wasslightly more effective thanpressure (P0.5=14 mmHg) inactivating MscCa

Pflugers Arch - Eur J Physiol (2006) 453:333–351 337

Page 6: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

selectivity target specific MS channels. These agents wouldbe highly useful for following MS channel proteins duringpurification procedures and identifying MS channel roles innovel functions. In addition, given that MS channels maybe polymodally activated (e.g., by tension, voltage, pH,temperature, Ca2+ store depletion, and/or lipid secondmessenger) [27, 125, 171], it would be advantageous todiscover agents that selectively acted on MS channel onlywhen they were mechano-gated. Although no ideal MSchannel reagent has yet been discovered, a number ofcompounds have been identified that act as MS channelblockers or activators [48, 61, 65]. One class, amiloride andits analogs, appear to act on a traditional “lock and key”protein receptor, whereas other agents, GsMTx-4 andpossibly maitotoxin, seem to act via nontraditional “recep-tors” at the lipid or lipid–protein interface where they may

change the local bilayer mechanics and thereby modify MSchannel gating. Below we briefly review their salientfeatures.

Amiloride has been the most rigorously studied in termsof its MS channel blocking mechanism and provides anexample where variations in mechanistic detail may enablediscrimination between different channel families in termsof their participation in specific MS functions. In particular,the amiloride block of MscCa/TRPC-1 in Xenopus oocytes[56, 89–91] and MscCa in vertebrate hair cells [141] hasbeen shown to involve basically the same unusual mech-anism in which two amiloride molecules bind cooperativelyto channel sites that only become accessible at hyper-polarized potentials after the channel has opened. Thismechanism, referred to as “conformational” block, impliesdifferent open-state conformations at hyperpolarized vs

Fig. 5 Changes in the membrane and patch currents as a consequenceof repetitive pressure pulses applied to the patch pipette. a–d videoimages of a cell-attached patch at different times after formation of thegiga-seal. Between each image, suction steps (20 mmHg, 500 ms)were applied. a The first image taken immediately after giga-sealformation shows the patch curved outwards and located closed to theend of the pipette (∼5 μm). Particles located in the cytoplasmexhibited no motion presumably because they were still constrainedby the intact CSK. b–d repetitive suction pulse caused the patch to

move up the pipette away from the cell, and a clear space developedbetween the membrane and the CSK remaining close to the cell.Particles that moved into this space displayed Brownian motion,indicating the loss of constraining CSK structures [153]. e Applicationof suction pulses at a caused a rapid opening of MscCa that closedalmost completely. f Application of a suction pulse at d causedactivation of a smaller more sustained current (from Zhang et al. [185]with permission)

338 Pflugers Arch - Eur J Physiol (2006) 453:333–351

Page 7: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

depolarized potentials, and is distinctly different from theamiloride block of the high-affinity amiloride-sensitiveepithelial Na+ channel (ENaC) where the voltage depen-dency arises because the positively charged amiloride binds

to a pore site that senses a fraction of the electric field[130]. A further difference is seen in the order of potenciesof amiloride analogs in blocking the two channel classes(Table 1) [80]. Amiloride blocks the MEC-4 DEG/ENaCcurrents in touch receptor neurons in Caenorhabditiselegans [120] and also the mammalian arterial myogenicresponse, which has been used to implicate DEG/ENaC asthe vascular mechanosensor [34]. However, amiloride alsoblocks TRPC-6, also implicated as the arterial smoothmuscle mechanosensor [74, 175]. It will be interesting todetermine if mechanistic differences in amiloride block(i.e., conformational vs pore block) can be used to resolvethe MS Channel_s identity.

Gadolinium blocks a variety of MS channels (e.g.,MscCa, MscK, MscL, and MscS), several TRP channels(e.g., TRPC-1, TRPC-4, TRPC-5, and TRPV1), variousvoltage-gated (Ca2+, Na+, and K+) and receptor-gatedchannels [e.g., N-methyl D-aspartate (NMDA), AChR,etc.] [61, 124, 136, 144, 165, 176, 178]. Because of itstrivalency, Gd3+ will bind with high affinity to negativegroups on proteins, lipids, and polysaccharides, as well asany inorganic anions present in solution [17, 61]. Its ionicradius (0.938 Å), which is similar to Na+ (0.97 Å) and Ca2+

(0.99 Å), may also allow it to enter and bind to negativelychange groups (Glu and/or Asp) within cation channels.Evidence that Gd3+ interacts directly with channel proteinscomes from studies of specific TRP members (TRPV-1,TRPC-4, and TRPC-5) where Gd3+ has been shown to havedual effects, activating the channels at low micromolarconcentrations (<100 μM) but blocking at higher concen-trations (>300 μM). The activation of TRPV-1 dependsupon binding to specific external glutamate residues thatconfer acid sensitivity on the channel, and neutralization ofthese residues blocks the activation and modifies inhibition[165]. Similar concentration-dependent potentiating andblocking effects also occur with TRPC-4 and TRPC-5[77]. In contrast, Gd3+ only blocks TRPC-1 and TRPC-3channels at relatively low micromolar concentrations [97,166].

0 20 40 60 80 100

mmHg

0

0.5

1.0

1.0

0.5

0

0

0.5

1.0

MscL

79 mmHg

MscK

40 mmHg

MscCa

10 mmHg

Fig. 6 Comparison of the normalized pressure–current relations forthree different MS channels. MscCa curve fitted to curve Fig. 2c (top).MscK curve fitted to data from Ref. [69] (middle). MscL curve fittedto data from Ref. [54] (bottom)

Table 1 Amiloride analog potency (IC50 amiloride/IC50 analog) of MS channels and the epithelial Na+ channel

Amiloride(IC50, μM)

DMA Phenamil PBDCB Benzamil HMA I-NMBA Reference no.

MscCa mouse hair cell 1 (53) 1.3 4.4 5 9.6 12.3 29 [141]MscCa Xenopus oocyte 1 (500) 1.4 – – 5.3 14.7 (BrHMA) [90]Epithelial Na+ channel(high affinity)

1 (0.34) 0.04 17 – 9 0.04 7 [80]

Epithelial Na+ channel(low affinity)

1 (10) 2.2 2.9 – 3.8 – – [80]

DMA Dimethylamiloride; PBDCB 5-(N-propyl-N-butyl)-dichlorobenzamile; HMA hexamethyleneamiloride; I-NMBA 6-iodide-2-methoxy-5-nitrobenzamile; BrHMA bromohexamethylene amiloride

Pflugers Arch - Eur J Physiol (2006) 453:333–351 339

Page 8: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

An early view on how Gd3+ might block MS channelswas via effects on the bilayer. Gd3+ has been shown tointeract with black lipid membranes containing the nega-tively charged phosphatidylserine (PS) but not with theneutral phosphatidylcholine to increase the boundarypotential and membrane tension [38]. However, whetherthese effects underlie MS channel block remains unclearbecause PS is normally restricted to the internal-facingmonolayer, and Gd3+ acts externally. Gd3+ has also beenshown to promote shape changes in giant unilamellarvesicles lacking PS [162]. In this case, it was proposedthat Gd3+ bound to the hydrophilic lipid head (i.e., negativecharge of the phosphate groups) of the external monolayer,and in doing so, decreased its surface area relative to theinternal monolayer, thereby causing a change in mem-brane curvature. However, whether this effect wouldblock MS channels remains unclear because amphipathsthat also change membrane curvature usually result in MSchannel activation [101, 135]. A recent study has reportedthat Gd3+ can block MS channels without alteringpressure-induced changes in Cm, which would beexpected if Gd3+ acted by altering membrane mechanics[156]. However, as pointed out by the authors, measure-ment of this parameter may be complicated because Gd3+

has multiple effects, including increasing the giga-seal[35].

GsMTx-4 is a 34-amino acid (4 kDa) peptide isolatedfrom tarantula venom [48, 49, 155, 157]. It is amphipathicwith a hydrophobic membrane face opposite a positivelycharged face, and it is a member of the inhibitory cysteineknot (ICK) toxin superfamily. GsMTx-4 is the most specificMS blocker identified to date. Because of its structure, itwould be expected to be attracted to negative regions ofproteins/lipids, and it will tend to partition into hydrophobicpockets either within the protein or at the protein/lipidinterface. Unlike the nonspecific channel blocker Gd3+,GsMTx-4 has so far not been reported to affect voltage- orreceptor-gated channel. GsMTx-4 blocks MscCa at between0.2 and 3 μM in chick heart, rat astrocyte, and humanbladder and kidney cells [48, 49], and the crude tarantulavenom also blocks MscCa in growing pollen protoplasts[36]. Most recently, GsMTx-4 has been shown to stimulateneurite outgrowth by blocking Ca2+ elevation in Xenopusspinal neurons [76]. However, GsMTx-4 does not blockMscCa involved in auditory transduction [48], MscKformed by TREK (E. Honore, unpublished observations,cited in Ref. [48]) the bacterial MscL [99], and perhaps,most surprisingly, MscCa in Xenopus oocytes (R. Marotoand O.P. Hamill, unpublished observations). The last resultis puzzling given that the oocyte MscCa is often treated asthe prototypical MscCa, and TRPC-1 has been implicatedas forming MscCa in the Xenopus oocyte and mammaliancells [97]. One possible explanation is that there are

structural differences between MscCa proteins in differentcell types. However, based on the observation that GsMTx-4synthesized from D instead of L amino acids shows the samepotency in blocking specific MS channels, it has beenproposed that GsMTx-4 is more likely to act by binding toboundary lipids surrounding the channel, and, at least,consistent with this is that GsMTx-4 and its enantomer alsoalters the gating of gramicidin A, which is particularlysensitive to bilayer mechanics [157]. It may therefore be thatdifferences in lipids between poikilotherms vs homeothermsis a factor that underlies the different GsMTx-4 sensitivities.In this case, it will be particularly interesting to determineGsMTx-4 action on MscCa/TRPCs reconstituted into de-fined lipid environments.

Maitotoxin (MTX) is a highly potent marine poison (LD50 for mice 50 ng/kg) from the dinoflagellate (Gambier-discus toxicus) that is responsible for Giguartera, a formof seafood poisoning [40]. It is water-soluble polyetherwith 2 sulfate esters, 28 hydroxyls, and 32 ether rings, andwith a molecular weight of 3.4 kDa, it is the largestamong the known nonbiopolymers. The hydroxyl andionized sulfate groups makes MTX a highly polarsubstance, but the presence of large hydrophobic portionsmake it amphipathic so that it most likely inserts itselfdeep into the bilayer. MTX elicits Ca2+ influx in a varietyof cell types, and the Ca2+ influx may lead to secondaryeffects, including phosphinositide breakdown and arachi-donic release. Of special interest here in that cellsexpressing TRPC-1 show a substantial increase in MTX-initiated Ca2+ influx that is blocked by Gd3+ (KD50=3 μM)and also by amiloride and benzamil but not by flufenamicacid or niflumic acid [10, 14, 174]. MTX activates 40 pSchannels when applied to outside-out patches butnot inside-out patches indicate that MTX acts on theextracellular face and does not require second messengers[40]. Both the conductance and pharmacologicalproperties have led to the idea that MTX activates theMscCa channels in oocytes, which is consistent withits effect of activating similar channel currents inTRPC-1 expressing cells. On the other hand, MTX alsoincreases Ca2+ influx in red blood cell (RBC) ghostswhich may involve another mechanism [85]. Although ithas been suggested that MTX mainly acts to increasecurrent by stimulating insertion of channels in the oocytemembrane, the evidence is based on large rapid Cm

changes that follow moment-to-moment changes inconductance induced by MTX and which are blocked bythe same ions and agents that also block the conductancechanges [174]. These properties indicate the Cm changesmay have been contaminated by changes in membraneconductance [25]. In this case, alternative methods formeasuring membrane trafficking (e.g., FMI-43 fluores-cence) should be used to test whether MTX-induced

340 Pflugers Arch - Eur J Physiol (2006) 453:333–351

Page 9: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

membrane conductance occurs by channel insertion and/or channel activation [40].

Mechanosensitive channel protein identification

The membrane proteins forming specific MS channelshave only been recently identified, and there were severalreasons for this delay, including the general low abun-dance of MS channels in animal cells, the absence ofhigh-affinity MS channel agents, the inability to employconventional expression cloning strategies because ofwidespread endogenous MS channel expression, and theabsence of identified mutant phenotypes involvingstretch-activated channels. To overcome these handicaps,a novel strategy was developed by Sukharev et al. [159,160] that involved detergent-solubilizing and fractionatingmembrane proteins, reconstituting the protein fractions inliposomes, then assaying the fractions for stretch sensitiv-ity using patch clamp recording. This technique has beenused to identify a variety of MS channel proteins inbacteria and archaea [81, 103, 158, 159] and, mostrecently, the TRPC protein family in forming MscCa inXenopus oocytes [97]. Furthermore, by demonstrating thata purified protein reconstituted in pure liposomes can retainstretch sensitivity, the technique also provided unequivocalevidence for the idea that bilayer tension alone gated MSchannels [101]. Although ENaC has also been reconsti-tuted in lipid bilayer and reported to show stretchsensitivity, it is not clear whether the proposed mechanismof stretch-induced release of Ca2+ channel block operatesin situ [75]. At this time, the best evidence for ENaC familymembers forming MS channels comes from genetic studiesof C. elegans touch-insensitive mutants [see 8, 39, 47, 62for reviews].

The proteins forming the other major class of MSchannels in animals cells, MscK [79, 118, 150, 168], wereidentified serendipitously, in that after the first membersof the 2 pore domain K+ (K2P) channel family hadalready been cloned and shown to form K+ channels [41,42, 93], they were subsequently found to be stretch-activated [6, 126, 129]. The recent demonstration thatTREK and TRAAK retain stretch sensitivity in CSK-freemembrane blebs indicates that they are also bilayer-gatedchannels [69]. In addition to the MS channel proteinsthat may function as mechanosensors in situ, there isalso an increasing number of voltage-gated and receptor-gated channels as well as peptides that form simplemodel channels (alamethicin and gramicidin) that displaymechanosensitivity [19, 102, 115, 122, 163]. Althoughthese channels may operate on the same general princi-ples that confer mechanosensitivity on membrane pro-teins, their role, if any, as mechanotransducers remains tobe demonstrated.

Mechanosensitive channel dynamics: adaptation/desensitization/inactivation

Gating dynamics (adaptation/inactivation/desensitization)has been shown to play a critical role in the signaling byvoltage- and receptor-gated channels and the hair cellmechanotransduction channel [67, 72] and defects in gatingdynamics underlie a number of channelopathies [4]. In theinitial studies of single MS channel currents, the channelsseemed to obey stationary kinetics and were analyzedaccordingly [46, 50, 51, 90, 114, 143, 177, 179]. However,with the ability to apply fast pressure steps to the patch [9,59, 105–108], it became evident that MS channels alsodisplayed dynamics in which the channels either closedreversibly (adaptation or inactivation/desensitization) orfaded irreversibly (run down) with constant stimuli [55,69, 105, 106, 156].

In principle, the reversible closure of MS channelsduring maintained stimulation can arise through relaxationin either the mechanical force being applied to the channelor the sensitivity to that mechanical force [54, 57] Becausemechanical gating arises from the channel protein beingsensitive to some mechanical-induced deformation [i.e.,either in the bilayer or in CSK/extracellular matrix (ECM)elements], then closure can arise because of a relaxation inthe force causing the deformation or a relaxation in thesensitivity to that deformation. For example, in the simplestcase of a two-state channel in which the rate constants forchannel opening (β) and closing (α) are displacement-sensitive (i.e., for a tethered MS channel) or tension-sensitive (i.e., for bilayer-gated MS channel) the probabilityof the channel being open (Po) will be given by:

Po ¼ 1 1þ Kð Þ=

where

K ¼ β a=

Or, in terms of displacement,

K ¼ K0es x0�xð Þ

where K0 is the equilibrium constant when the displacementx is equal to the set point x0 and determines the number ofchannels open at zero relative displacement, and s is thesensitivity to the relative displacement change (x0−x). For abilayer-gated channel, we can substitute displacement witharea change. An exponential time relaxation in either s or x0can produce the same adapting MS channel currents [57].Figure 7 illustrates simulations of the stimulus-responserelations made, assuming that after a step stimulus, there isan exponential change in either the set point x0 (Fig. 7a) orthe sensitivity factor s (Fig. 7b).

Although both mechanisms predict the same kinetics ofchannel closure, the consequences on the Po−X curves are

Pflugers Arch - Eur J Physiol (2006) 453:333–351 341

Page 10: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

clearly different. In the first case of adaptation, the curvesshift along the x-axis with no changes in slope (i.e.,sensitivity) around a common set point. In the second case,

the sensitivity decreases as the curves pivot around acommon point. From a functional point of view, the firstcase is true adaptation because sensitivity is maintained

Fig. 7 Simulation of two mechanisms that results in closure of MSchannels in the presence of sustained stimulation. A two-state channelis assumed, and a step displacement from 0 to x (top, trace 1) is usedto activate the MS channels (a and b). Trace 2 represents the changesin tension (a) or sensitivity tension (b). In trace 3, the channel currentsare represented by the change in open-channel probability (Po), withthe numbered points (1–5) representing equally spaced times wherePo–X curves were generated to follow changes in the MS channelsensitivity. a The decay of the current is due to a change in the tension

(measured as x−x0) caused by a shift in the set point. In this case, thereis a shift along the x-axis with no change in slope. Consequently, theδPo response due to δx does not decrease during what can beconsidered true adaptation. b The decay of the current is due to achange in sensitivity in which the slopes of the Po–X curves decreaseas they pivot around a common point. As a consequence, theincremental change in the response (δPo) for a fixed δx decreasesduring the current decay, which is akin to receptor desensitization.Modified from Ref. [57]

342 Pflugers Arch - Eur J Physiol (2006) 453:333–351

Page 11: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

[55], whereas the second case is more akin to receptordesensitization or voltage-gated channel inactivation, wherethe stimulus must be removed for sensitivity to recover[69]. Below, we consider specific MS channels in terms ofthese general principles.

MscCa in Xenopus oocytes This channel displays differentgating dynamics depending upon patch “history.” In thecase when the giga-seal is formed using a “gentle” suctionprotocol (e.g., 10 mmHg for 1–10 s), the application of asuction/pressure step produces rapid opening (<1 ms) ofchannels followed by a slower closure (∼100 ms), althoughthe stimulus is maintained constant. The resultant decay ofMscCa current can be fitted by a single exponential with atime constant of around 100 ms at −100 mV that shows amonotonic e-fold increase for every approximately 100-mVdepolarization. The voltage dependence of the channelclosure is most evident when the voltage is switched fromhyperpolarized to depolarized potential (or vice versa)during the pressure step (Fig. 8) [55, 107, 108]. Thedirection of this voltage dependence is similar to thevoltage dependence of adaptation displayed by the hair cellmechanotransducer channel [5] and MscS (see below). Thestimulus induced closure of the oocyte channel wasoriginally referred to as adaptation because increasing thestimulus could reopen channels. In the oocyte, suctions/pressures of approximately 20 mmHg produce saturatingresponses (see Fig. 4), so it was assumed that any channelopening caused by an increase in suction/pressure of at least20 mmHg would involve reopening channels that had justclosed. However, a practical limitation in using theseprotocols on oocyte patches is that application of evenlarger pulses (e.g., ≥40 mmHg) that would undoubtedlyactivate all channels will also cause irreversible loss of thechannel activity and gating dynamics as described below[55, 105, 106].

In the second case, if a more forceful suction protocol isrequired to achieve the seal, then more often than not, thetransient current response is absent, and instead, thechannels remain open for the full duration of the suction.Similarly, if after a gentle seal is formed the patch ismechanically “over-stimulated,” then adaptation of MscCaactivity disappears either progressively with each moderate-sized pulse (Fig. 9a) or suddenly within a single largesuction (Fig. 9b). This transition from the transient mode(TM) to the sustained mode (SM) of gating is irreversibleand occurs without a change in single-channel conductance[64].

The fragility of MscCa dynamics and the transition fromTM to SM gating has been proposed to arise throughmechanical decoupling of CSK interactions with either thechannel or the membrane, which are thought to beimportant for TM of gating. It has been suggested that

viscous elements (dashpots) in the CSK can become frozenor decoupled without disconnecting the gating springs.However, adaptation is preserved in both inside-out andoutside-out patches, and in patches treated with agents thatdisrupt microtubules (colchicine) or microfilaments (cyto-chalasin D), similar to what has been reported for TRAAKdesensitization [69]. In contrast, transient gating kinetics ofMscCa are not retained in either “blebbed” membrane thatit lacks an underlying CSK [185] or in pure liposomepatches expressing MscCa activity following reconstitutionof detergent-solubilized oocyte membrane proteins [97].Furthermore, overexpression of TRPC-1 that forms theoocyte MscCa does not result in channel activity thatdisplays TM gating. Whether the absence of adaptationreflects prior mechanodisruption or the absence of CSKremains unclear, as does the mechanism that causesirreversible run down. One possibility is that there are

Fig. 8 Voltage dependence of pressure-induced currents recordedfrom cell-attached oocyte patches. In each panel, the top tracerepresents the pressure (suction); the middle trace, the voltage; andthe bottom trace, the current. a The application of the suction pulse tothe patch held at −100 mV caused rapid opening of channels that hadnearly all closed before the voltage was switched to 100 mV whilemaintaining the suction pulse. b In this case, the suction was appliedto the patch held 100 mV and produced a steady-state current.Switching the voltage to −100 mV activated a transient increase incurrent that decayed incompletely in the presence of maintainedsuction

Pflugers Arch - Eur J Physiol (2006) 453:333–351 343

Page 12: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

irreversible changes in the membrane–glass adhesion thatalters the ability to generate tension changes in the bilayer.For example, if the membrane does not reseal aftermechanical decoupling of lipid–glass interface [121], thenincreasing pressure may draw further membrane into thepipette without increasing bilayer tension.

MscCa in rat astocytes This channel shows certain dynamicproperties similar to those of the oocyte MscCa, including itsvoltage dependence and mechanical fragility. However, inthe astrocyte, the decrease in current occurs because ofincreased occupancy of lower conductance states and areduced open-channel probability [13, 156]. Furthermore,the closed channels cannot be reactivated by increasing thestimulus strength (i.e., are refractory), indicating inactiva-tion rather than adaptation. The process was modeled as aball-and-chain-type inactivation, in which the inactivatingball was a CSK element rather than part of the channelprotein. By assuming that the binding rates of theinactivating ball were affected by the position of anintramembrane voltage-sensing subunit, one can accountfor the voltage dependence of inactivation. Apparentlyconsistent with the model, it was demonstrated that a

combination of agents targeting actin (cytochalasin), micro-tubules (colchicine), and intermediate filaments (acrylam-ide) was required to abolish the inactivation, but this lossmight again reflect general mechanical patch damage ratherthan implicating specific CSK elements. In the same study,fast Cm measurements were used to monitor the changein membrane area/thickness during pressure steps anddemonstrated a similar voltage-independent monotonicincrease in patch capacitance at −90 and + 50 mV, whichcontrasted with the voltage-dependent inactivation. Thisobservation was interpreted as indicating inactivation was dueto intrinsic properties of the channel rather than relaxation ofbilayer tension [156].

MscS and MscL in Escherichia coli The two predominantMS channels in Escherichia coli, MscS (0.5 pS) and MscL(1–3 nS) [100, 160], also exhibit transient gating dynamics[2, 66, 86, 87]. MscS currents measured in E. coliprotoplasts in response to pressure steps undergo apressure-induced exponential decay that appeared to bevoltage-independent with a time constant of 2–3 s whenmeasured over a narrow voltage range of ±30 mV [86].However, when measured over a wider voltage range of

Fig. 9 Irreversible loss of tran-sient mode gating of MscCa.a Three consecutive suctionpulses of 30 mmHg were ap-plied to a patch 30 s apart,causing a progressive loss of thetransient gating and a decreasein the peak current. b A singlelarge suction pulse (100 mmHg,10 s) was applied to a patch andcaused an initial peak currentthat was converted into a sus-tained current

344 Pflugers Arch - Eur J Physiol (2006) 453:333–351

Page 13: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

±100 mV there was a decrease in the rate of inactivationwith depolarization similar to MscCa [2]. Application of thesecond of a double-step protocol activates fewer channels,and a finite time without stimulation is required forrecovery of the full response [86]. Furthermore, suctionramps produced smaller peak responses than suction steps[2]. These features are more consistent with inactivationthan adaptation. Unlike MscCa in vertebrate cells, MscSdynamics are not mechanically fragile, although pronasetreatment of the intracellular membrane face abolishes thetransient kinetics and, ultimately, mechanosensitivity. Thislast observation led Koprowski and Kubalski to proposethat both activation and inactivation may depend uponinteraction between a cytoplasmic (pronase-sensitive) re-gion of the channel with the lipid bilayer [86, 87]. Note theproteolytic inhibition of MscS activity is opposite to thepotentiation of MscL activity [1]. Given that a bilayer ratherthan a tethered mechanism gates MscS, it was proposed thatinactivation might be associated with insertion of thecytoplasmic domain of MscS in the bilayer (i.e., a “hybrid”or intrinsic tethered model; see below).

MscL reconstituted in liposomes also shows a transientdecay in the current with a time constant of seconds [66].Although the distinction between adaptation and inactivationstill needs to be made, the observation is significant becausethe clear absence of any CSK excludes its involvement inthese transient kinetics. One possible explanation is time-dependent sliding/relaxation of the two monolayers thatresults in relaxation of the gating tension [144].

TREK and TRAAK MS channels In a recent study, pressuresteps have been used to analyze the dynamics of MscKformed by cloned TREK-1 and TRAAK herologouslyexpressed in COS cells and Xenopus oocytes [69]. Bothchannels show rapid closure (τ∼20–50 ms), with constantstimulation similar to the MscCa. However, unlike MscCa,MscK gating dynamics are not voltage-sensitive, and eithermechanical or chemical disruption (i.., using latrunculin) ofthe CSK causes “run up” rather that “run down” of thechannels without removing the transient gating kinetics.Because it was clearly demonstrated that channels could notbe reactivated without a finite time for recovery, thephenomenon was referred to as desensitization. The lack

Fig. 10 Three different modelsof mechanosensitive channelgating a Bilayer. Mechanicalforces are conveyed to thechannel purely via the bilayer.Tension sensitivity occurs be-cause of a difference proteinarea (or hydrophobic thicknessand/or lateral shape) betweenthe open and closed channelconformations. b Extrinsic teth-er. Tensions are exerted directlyon the channel protein via ex-tracellular or cytoskeletal elasticelements/gating springs. Whentension is exerted on the gatingspring, the open state is ener-getically more favorable. Intrin-sic tether (hybrid). In this model,the gating spring is one of thecytoplasmic domains that bindsto the phospholipids and, in thisway, becomes sensitive tomembrane stretch

Pflugers Arch - Eur J Physiol (2006) 453:333–351 345

Page 14: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

of effects of either mechanical or chemical CSK disruptionindicates that desensitization is an intrinsic property of thechannel [69].

Molecular models of stretch sensitivity

There are three broad classes of mechanisms that mayimpart stretch sensitivity on a membrane ion channel.They will be referred to as “bilayer,” “tethered,” and“hybrid,” and are shown schematically in Fig. 10. Themodels need not be mutually exclusive, and a singlechannel may derive its mechanosensitivity from all threemechanisms. Each mechanism can be discussed in termsof a simple two-state channel that fluctuates between aclosed and open state. The bilayer mechanism applies to avariety of MS channels, as evidenced by retention ofmechanosensitivity following liposome reconstitutionand/or activation by amphipaths or lysophospholipids.The basic idea is that stretching the bilayer will tend todecrease its lipid packing density and thickness, so that ifthe channel protein undergoes a change in membrane-occupied area (Fig. 9a) and/or hydrophobic mismatch,there will be a shift in the distribution between closed andopen channel conformations [54, 88, 95]. By inserting inthe membrane, lysophospholipids and amphipathic mole-cules may cause local changes in tension and curvature atthe lipid–protein interface and thereby shift the channeldistribution [12, 95, 98, 99, 131–133].

In the tethered mechanism, either an extracellular orcytoskeletal protein is directly connected to the channeland acts as a gating spring [50, 58, 62, 71, 72]. When thegating spring is stretched, it favors the open state of thechannel because it allows relaxation of the spring. In

Fig. 9b, the gate is represented as a trapdoor that opensout, but it can well represent subunits that are eitherpulled apart (increased in area) or lengthened (change ifhydrophobic mismatch). Evidence pro and con for thetethered mechanism has been discussed previously [54,88].

The hybrid of the above two mechanisms depends uponstretching of the bilayer, but in this case, there arecytoplasmic domains of the channel protein that bind tophospholipids, and in this way act as intrinsic tethers orgating springs that are stretched along with the bilayer(Fig. 9c). Evidence for the hybrid model comes from theidentification in the specific K2P channels of a phospho-lipid-sensing domain on the proximal carboxyl terminusthat involves a cluster of positively charged residues thatalso includes the proton sensor E306 [23, 68]. Protonationof E306 drastically tightens channel–phospholipid interac-tion and leads to TEK-1 opening at atmospheric pressure.The carboxy terminal domain of TREK-1 interacts withplasma membrane, probably via electrostatic interactionbetween a cluster of positive charges (a PIP2-interactingdomain) and anionic phospholipids.

Mechanosensitive channels in human diseases

An exiting development in the field has been the growingnumber of diseases associated with abnormalities ofmechanotransduction. Donald Ingber [73], in a recentreview, listed 45 diseases that may arise due to changes incell mechanics, alterations in tissue structure, or deregula-tion of mechanosignaling pathways. Of these diseases,several have been directly associated with changes inexpression and/or gating of MS channels, including cardiacarrhythmias [84], polycystic kidney disease [18], hyperten-

Fig. 11 Cell-attached patch re-cording on an mdx mouse myo-tube showing high constitutivechannel activity that was re-duced with suction but unaf-fected by positive pressure. This“apparent” stretch-inactivatedchannel activity was rare andseen in only 1 of approximately100 patches. The majority ofother patches showed no spon-taneous channel activity, andsuction activated either a tran-sient opening of channels orchannels that remained open for10 s after the pulse (e.g., see[107])

346 Pflugers Arch - Eur J Physiol (2006) 453:333–351

Page 15: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

sion [83], glioma [123], glaucoma [78] atherosclerosis [22,134], Duchenne muscular dystrophy [44, 45], and tumor-igenesis [128]. Furthermore, increased MscK activity hasbeen shown to prevent brain ischemia [16] and promotegeneral anesthesia [127], whereas MscCa/TRPC activitymay regulate wound healing [137] and promote neuronalregeneration [76]. Of particular note is Duchenne musculardystrophy (DMD), a devastating X-linked genetic diseasethat affects approximately 1 in 3,500 male births and ischaracterized by progressive muscle wasting and weakness(reviewed in [180]). DMD is caused by the absence of thegene product of dystrophin, a cytoskeletal protein that bindsto actin and provides structural support for the membraneparticularly during muscle stretching. In mdx muscle fibers(i.e., from the mouse model of DMD), there is increasedvulnerability to stretch-induced membrane wounding, andseveral studies indicate elevated [Ca2+]i levels in mdxmyotubes that have been associated with increased Ca2+

permeant leak channel activity [43] and/or abnormal MscCaactivity [44, 45]. Anti-MscCa agents, including Gd3+,streptomycin, amiloride, and GsMTx-4, have been reportedto block Ca2+ elevation and/or reduce muscle fiberdegeneration [3, 56]. Based on the observation that theleak channel activity was increased by internal calciumstore depletion, Vandebrouck et al. [167] proposed that astore-operated Ca2+ channel (SOCC) belonging to theTRPC family may be involved. To test this idea, theytransfected muscles with antisense oligonucleotide designedagainst the most conserved region sequences of the TRPCsand showed it caused significant knockdown of TRPC-1and - 4 but not TRPC-6 (all three were detected in wild-type and mdx muscle), and reduced both control andthapsigagin-induced Ca2+ leak channels without affectingvoltage-gated Na+ channels. The mechanosensitivity of thechannels was not tested in this study. However, MscCa canshow significant spontaneous opening in the absence ofmembrane stretch [140]. Furthermore, although Franco andLansman [44] initially reported a stretch-inactivated Ca2+

channel in mdx mice, they subsequently concluded that thechannel activity may arise from a novel gating mode ofMscCa induced by membrane stress [45]. Most recently, ithas been suggested that stretch inactivation in patches ofmdx muscle and other cells may be a patch recordingartifact induced when suction applied to the patch reduces atonic tension generated by CSK forces that bend the patchtoward the cell [69]. At least consistent with this notion isthat suction (but not positive pressure) causes inactivationof MscCa in mdx patches (Fig. 11).

Conclusion and future prospects

The giga-seal patch clamp technique has been a majorcontributor to increased understanding of MS channels over

the last 20 odd years. However, there is still somewhat adisconnect between the phenomena seen in the patch andhow they translate in MS currents in the whole cell.Furthermore, given the growing evidence that MS channelsare promiscuous in terms of their modes of activation, itbecomes even more important to identify the exactphysiological stimulus that activates the channel in specificsituations. The development of new techniques that canmonitor/generate membrane tension changes in normallyoperating cells while recording MS channel on the cell canaddress many of the unresolved issues. Similarly, thediscovery of high affinity and selective agents that cantarget mechanically gated channels will represent a majorbreakthrough for the field. The determination of the crystalstructure of bacterial MS channels [7, 21, 139] has provideda rich environment for model building and testing, and asimilar trajectory is predicted for the recently identified MSchannel proteins in animal cells. A key question that thesestudies should answer is whether a unified set of principlescan account for the stretch sensitivity of channels in bothprokaryotes and eukaryotes [54, 88].

Acknowledgments Research in the author’s laboratory is funded bythe National Institutes of Health (NIH) and the Department of Defense(DOD).

References

1. Ajouz B, Berrier C, Besnard M, Martinac B, Ghaz A (2000)Contributions of the different extramembranous domains of themechanosensitive ion channel MscL to its response to membranetension. J Biol Chem 275:1015–1022

2. Akitake B, Anishkin A, Sukharev S (2005) The “dash-pot”mechanism of stretch-dependent gating in MscS. J Gen Physiol125:143–154

3. Allen DG, Whitehead NP, Yeung EW (2005) Mechanisms ofstretch-induced muscle damage in normal and dystrophicmuscle: role of ionic changes. J Physiol 567.3:723–735

4. Ashcroft FM (2006) From molecule to malady. Nature 440:440–447

5. Assad JA, Hacohen N, Corey DP (1991) Voltage dependence ofadaptation and active bundle movement in bull frog saccular haircells. Proc Natl Acad Sci USA 86:2918–2922

6. Bang H, Kim Y, Kim D (2000) TREK-2, a new member of themechanosensitive tandem-pore K+ channel family. J Biol Chem275:17412–17419

7. Bass RB, Strop P, Barclay M, Rees DC (2002) Crystal structureof Escherichia coli MscS, a voltage-modulated and mechano-sensitive channel. Science 298:1582–1592

8. Bazopoulou D, Tavernarakis N (2007) Mechanosensitive ionchannels in Caenorhabditis elegans. In: Hamill OP (ed)Mechanosensitive channels. Elsevier, Amsterdam (in press)

9. Besch SR, Suchyna T, Sachs F (2002) High speed pressureclamp. Pflugers Achiv 445:161–166

Pflugers Arch - Eur J Physiol (2006) 453:333–351 347

Page 16: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

10. Bielfeld-Ackermann A, Range C, Korbmacher C (1998) Maito-toxin (MTX) activates a nonselective cation channel in Xenopuslaevis oocytes. Pflugers Archiv 436:329–337

11. Bittner GD, Fishman HM (2000) Axonal sealing followinginjury. In: Ingola N, Murray M (eds) Nerve regeneration. Dekker,New York, pp 337–370

12. Blount P (2003) Molecular mechanisms of mechanosensation:big lessons from small cells. Neuron 37:731–734

13. Bowman CB, Lohr JW (1996) Mechanotransducing ion channelsin C6 glioma cells. Glia 18:161–176

14. Brereton HM, Chen J, Rychkov G, Harland ML, Barritt GJ(2001) Maitotoxin activates an endogenous non-selective cationchannel and is an effective initiator of the activation of theheterologously expressed hTRPC1 (transient receptor potential)non-selective cation channel in H4-IIE liver cells. BiochimBiophys Acta 1540:107–126

15. Bryan-Sisneros AA, Fraser SP, Djamgoz MBA (2003) Electro-physiological mechanosensitive responses of Xenopus laevisoocytes to direct, isotonic increase in intracellular volume.J Neurosci 125:103–111

16. Buckler KJ, Honore E (2005) The lipid-activated two poredomain K+ channel TREK-1 is resistant to hypoxia: implicationfor ischaemic neuroprotection. J Physiol 562.1:213–222

17. Caldwell RA, Clemo HF, Baumgarten CM (1998) Usinggadolinium to identify stretch-activated channels: technicalconsiderations. Am J Physiol C619–C621

18. Cantiello HF (2003) A tale of two tails: ciliary mechanotrans-duction in ADPKD. Trends Mol Med 9:234–236

19. Casado M, Ascher P (1998) Opposite modulation of NMDAreceptors by lysophospholipids and arachidonic acid: commonfeatures with mechanosensitivity. J Physiol 513:317–330

20. Cemerikic D, Sackin H (1993) Substrate activation of mechano-sensitive, whole cell currents in renal proximal tubule. Am JPhysiol 264:F697–F714

21. Chang G, Spencer RH, Lee AT, Barclay MT, Rees DC (1998)Structure of the mscL homolog from Mycobacterium tubercu-losis: a gated mechanosensitive ion channel. Science 282:220–2226

22. Chapleau MW, Cunningham JT Sullivan MJ, Wachtel REAbboud FM (1995) Structural versus functional modulation ofthe arterial baroreflex. Hypertension 26:341–347

23. Chemin J, Patel AJ, Duprat F, Lauritzen I, Lazdunski M, HonoreE (2005) A phospholipid sensor controls mechanogating of theK+ channel TREK-1. EMBO J 24:44–53

24. Chen BM, Grinnell AD (1995) Integrins and modulation oftransmitter release from motor terminals by stretch. Science269:1578–1580

25. Chen P, Hwang TC, Gillis KD (2001) The relationship betweencAMP, Ca2+ and transport of CFTR to the plasma membrane. JGen Physiol 118:135–144

26. Christensen O (1987) Mediation of cell volume regulation byCa2+ influx through stretch-activated cation channels. Nature330:66–68

27. Clapham DE (2003) TRP channels as cellular sensors. Nature426:517–524

28. Dai J, Sheetz MP (1999) Membrane tether formation fromblebbing cells. Biophys J 77:3363–3370

29. Dai J, Sheetz MP (1995) Mechanical properties of neuronalgrowth cone membranes studied by tether formation with lasertweezers. Biophys J 68:988–996

30. Dai J, Sheetz MP, Wan X, Morris CE (1998) Membrane tensionin swelling and shrinking molluscan neurons. J Neurosci18:6681–6692

31. Dai J, Ting-Beall HP, Sheetz MP (1997) The secretion-coupledexocytosis correlates with membrane tension changes in RBL2H3 cells. J Gen Physiol 110:1–10

32. Disher DE, Mohandas N (1996) Kinematics of red cellaspiration by fluorescence-imaged microdeformation. BiophysJ 71:1680–1694

33. Discher DE, Mohandas N, Evans EA (1994) Molecular maps ofred cell deformation: hidden elasticity and “it situ” connectivity.Science 266:1032–1035

34. Drummond HA, Gebremedhim D, Harder DR (2004) Degenerin/Epithelial Na+ channel protein. Components of a vascularmechanosensor. Hypertension 44:643–648

35. Dunina-Barkovskaya AY, Levina NN, Lew RR, Heath IB (2004)Gadolinium effects on gigaseal formation and the adhesiveproperties of a fungal amoeboid cell, the slime mutant ofNeurospora crassa. J Membr Biol 198:77–87

36. Dutta R, Robinson KR (2004) identification and characterizationof stretch-activated ion channels in pollen protoplasts. PlantPhysiol 135:1398–1406

37. Erickson CA, Trinkaus JP (1976) Microvilli and blebs as sourcesof reserve surface membrane during cell spreading. Exp Cell Res99:375–384

38. Ermakov YA, Averbakh AZ, Yusipovich AI, Sukarev S (2001)Diploe potentials indicate restructuring of the membrane inter-face induced by gadolinium and beryllium ions. Biophys J80:1851–1862

39. Ernstrom GG, Chalfie M (2002) Genetics of sensory mechano-transduction. Ann Rev Genet 36:411–453

40. Escobar LI, Salvador C, Martinez M, Vaca L (1998) Maitotoxin,a cationic channel activator. Neurobiology 6:59–74

41. Fink M, Duprat F, Lesage F, Reyes R, Romey G, Heurteaux C,Lazdunski M (1996) Cloning, functional expression and brainlocalization of a novel unconventional outward rectifier K+

channel. EMBO J 15:6854–686242. Fink M, Lesage F, Duprat F, Heurteaux C, Reyes R, Fosset M,

Lazdunski M (1998) A neuronal two pore K+ channel activatedby arachidonic acid polyunsaturated fatty acids. EMBO J17:3297–3308

43. Fong P, Turner PR, Denetclaw WF, Steinhardt RA (1990)Increased activity of calcium leak channels in myotubes ofDuchenne human and mdx mouse origin. Science 250:673–676

44. Franco A, Lansman JB (1990) Calcium entry through stretch-inactivated channels in mdx myotubes. Nature 344:670–673

45. Franco-Obregon A, Lansman JB (2002) Changes in mechano-sensitive channel gating following mechanical stimulation inskeletal muscle myotubes from the mdx mouse. J Physiol539.2:391–407

46. Gil Z, Magleby KL, Siberberg SD (2001) Two-dimensionalkinetic analysis suggests nonsequential gating of mechanosensi-tive channels in Xenopus oocytes. Biophys J 81:2082–2089

47. Goodman MB, Schwarz EM (2003) Transducing touch inCaenorhabditis elegans. Ann Rev Physiol 65:429–452

48. Gottlieb PA, Suchyna TM, Ostrow LW, Sachs F (2004)Mechanosensitive ion channels as drug targets. Curr DrugTargets 3:287–295

49. Gottlieb PA, Suchyna TM, Sachs F (2007) Properties andmechanism of the mechanosensitive ion channel blockerGsMTx4, a therapeutic peptide derived from tarantula venom.In: Hamill OP (ed) Mechanosensitive channels. Elsevier,Amsterdam (in press)

50. Guharay F, Sachs F (1984) Stretch-activated single ion channelcurrents in tissue cultured embryonic chick skeletal muscle. JPhysiol 352:685–701

51. Guharay F, Sachs F (1985) Mechanotransducer ion channels inchick skeletal muscle: the effects of extracellular pH. J Physiol363:119–134

52. Hamill OP (1983) Potassium and chloride channels in red bloodcells. In: Sakmann B, Neher E (eds) Single channel recording.Plenum, New York, pp 451–471

348 Pflugers Arch - Eur J Physiol (2006) 453:333–351

Page 17: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

53. Hamill OP, Marty A, Neher E, Sakmann B, Sigworth F (1981)Improved patch clamp techniques for high current resolutionfrom cells and cell-free membrane patches. Pflugers Arch391:85–100

54. Hamill OP, Martinac B (2001) Molecular basis of mechano-transduction in living cells. Physiol Revs 81:685–740

55. Hamill OP, McBride DW Jr (1992) Rapid adaptation of themechanosensitive channel in Xenopus oocytes. Proc Natl AcadSci USA 89:7462–7466

56. Hamill OP, Lane JW, McBride DW Jr (1992) Amiloride: amolecular probe for mechanosensitive channels. Trends PharmacolSci 13:373–376

57. Hamill OP, McBride DW Jr (1994) Molecular mechanisms ofmechanoreceptor adaptation. News Physiol Sci 9:53–59

58. Hamill OP, McBride DW Jr (1994) The cloning of a mechano-gated membrane channel. Trends Neurosci 17:11:439–443

59. Hamill OP, McBride DW Jr (1995) Pressure/patch-clampmethods. In: Boulton AA, Baker GB, Walz W (eds) Patch clamptechniques and protocols. Humana, Totowa, pp 75–87

60. Hamill OP, McBride DW Jr (1995) Mechanoreceptive membraneion channels. Am Sci 83:30–37

61. Hamill OP, McBride DW Jr (1996) The pharmacology ofmechanogated membrane ion channels. Pharmacol Rev48:231–252

62. Hamill OP, McBride DW Jr (1996) A supramolecular complexunderlying touch sensitivity. Trends Neurosci 19:258–261

63. Hamill OP, McBride DW Jr (1997) Induced membrane hypo-/hyper-mechanosensitivity: a limitation of patch clamp recording.Annu Rev Physiol 59:621–631

64. Hamill OP, McBride DW Jr (1997) Mechanogated channels inXenopus oocytes: different gating modes enable a channel toswitch from a phasic to a tonic mechanotransducer. Biol Bull192:121–122

65. Hamill OP, McBride DW Jr (1998) Drug effects on mechano-gatedchannels. In: Soria B, Cena V (eds) The pharmacology of membraneion channel. Oxford University Press, Oxford, pp 271–278

66. Häse CC, Le Dain AC, Martinac B (1995) Purification andfunctional reconstitution of the recombinant large mechanosen-sitive ion channel (MscL) of Escherichia coli. J Biol Chem270:18329–18334

67. Hille B (2001) Ion channels of excitable membranes, 3rd edn.Sinauer, Sunderland, pp 1–814

68. Honoré E, Maingret F, Lazdunski M, Patel AJ (2002) Anintracellular proton sensor commands lipid and mechano-gatingof the K+ channel TREK-1. EMBO J 21:2968–2976

69. Honoré E, Patel AJ, Chemin J, Suchyna T, Sachs F (2006)Desensitization of mechano-gated K2P channels. Proc Natl AcadSci USA 103:6859–6864

70. Hörber JKH, Mosbacher J, Häbele W, Ruppersberg JP, SakmannB (1995) A look at membrane patches with scanning forcemicroscope. Biophys J 68:1687–1693

71. Howard J, Bechstedt S (2004) Hypothesis: a helix of ankyrinrepeats of the NOMPC–TRP ion channel is the gating spring ofmechanoreceptors. Curr Biol 14:224–226

72. Howard J, Roberts WM, Hudspeth AJ (1988) Mechanoelectricaltransduction by hair cells. Annu Rev Biophys Chem 17:99–124

73. Ingber DE (2003) Mechanobiology and diseases of mechano-transduction. Ann Med 35:1–14

74. Inoue R, Okada T, Onoue H, Hara Y, Shimizu S, Naitoh S,Ito Y, Mori Y (2001) The transient receptor potential proteinhomologue TRP6 is the essential component of vascular 1-adrenoceptor-activated Ca2+-pereable cation channel. Circ Res88:325–332

75. Ismailov II, Berdiev BK, Shlyonsky VG, Benos DJ (1997)Mechanosensitivity of an epithelial Na+ channel in planar lipidbilayers: release from Ca2+ block. Biophys J 72:1182–1192

76. Jacques-Fricke BT, Seow Y, Gottlieb PC, Sachs F, Gomez TM(2006) Ca2+ influx through mechanosensitive channels inhibitsneurite outgrowth in opposition to other influx pathways andrelease of intracellular stores. J Neurosci 26:5656–5664

77. Jung S, Mühle A, Shaefer M, Strotmann R, Schultz G, Plant TD(2003) Lanthanides potentiate TRPC5 currents by an action atthe extracellular sites close to the pore mouth. J Bio Chem278:3562–3571

78. Kalapesi FB, Tan JCH, Coroneo MT (2005) Stretch-activatedchannels; a mini-review: are stretch-activated channels an ocularbarometer? Clin Exp Opthal 33:210–217

79. Kim D (1992) A mechanosensitive K+ channel in heart cellsactivation by arachidonic acid. J Gen Physiol 100:1021–1040

80. Kleyman TR, Cragoe EK (1988) Amiloride and its analogs in thestudy of ion transport. J Membr Biol 105:1–21

81. Kloda A, Martinac B (2001) Mechanosensitive channels ofThermoplasma, the cell wall-less Archaea. Cell BiochemBiophys 34:321–347

82. Knutton S, Jackson D, Grahman JM, Micklem KJ, Pasternak CA(1976) Microvilli and cell swelling. Nature 262:52–53

83. Kohler R, Distler A, Hoyer J (1999) Increased mechanosensitivecurrents in aortic endothelial cells from genetically hypertensiverats. J Hypertens 17:365–371

84. Kohl P, Franz, MR, Sachs F (2002) Cardiac mechano-electricfeedback and arrhythmias: from pipette to patient. Elsevier,Amsterdam

85. Konoki K, Hashimoto M, Muata M, Tachibana K (1999)Maitotoxin-induced calcium influx in erythrocyte ghosts andrat glioma C6 cells and blockade by gangliosides and othermembrane lipids. Chem Res Toxicol 12:993–1001

86. Koprowski P, Kubalski A (1998) Voltage-independent adaptationof mechanosensitive channel in Escherichia coli protoplasts.J Membr Biol 164:253–262

87. Koprowski P, Kubalski A (2003) C termini of the Escherichiacoli mechanosensitive ion channel (MscS) move apart upon thechannel opening. J Biol Chem 278:11237–11245

88. Kung C (2005) A possible unifying principle for mechanosensa-tion. Nature 436:647–654

89. Lane JW, McBride DW Jr, Hamill OP (1993) Ionic interactionswith amiloride block of the mechanosensitive channel inXenopus oocytes. Br J Pharmacol 108:116–119

90. Lane JW, McBride DW Jr, Hamill OP (1992) Structure–activityrelations of amiloride and some of its analogues in blocking themechanosensitive channel in Xenopus oocytes. Br J Pharmacol106(2):283–286

91. Lane J, McBride DW Jr, Hamill OP (1991) Amiloride block ofthe mechanosensitive cation channel in Xenopus oocytes.J Physiol 441:347–366

92. Lee J, Ishihara A, Oxford G, Johnson B, Jacobson K (1999)Regulation of cell movement is mediated by stretch-activatedcalcium channels. Nature 400:382–386

93. Lesage F, Guillemare Fink M, Duprat F, Lazdunski M, RomeyG, Barhanin J (1996) TWIK-1, a ubiquitous human weaklyinward rectifying K+ channel with a novel structure. EMBO J15:1004–1011

94. Levina N, Tötemeyer S, Stokes NR, Louis P, Jones MA, BoothIR (1999) Protection of Escherichia coli cells against extremeturgor pressure by activation of MscS and MscL mechanosensi-tive channels: identification of genes for MscS activity. EMBO J18:1730–1737

95. Markin VS, Sachs F (2004) Thermodynamics of mechanosensi-tivity. Phys Biol 1:110–124

96. Maroto R, Hamill OP (2001) Brefeldin A block of integrin-dependent mechanosensitive ATP release from Xenopus oocytesreveals a novel mechanism of mechanotransduction. J BiolChem 276:23867–23872

Pflugers Arch - Eur J Physiol (2006) 453:333–351 349

Page 18: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

97. Maroto R, Raso A, Wood TG, Kurosky A, Martinac B, HamillOP (2005) TRPC1 forms the stretch-activated cation channel invertebrate cells. Nature Cell Biol 7:1443–1446

98. Martinac B (2004) Mechanosensitive ion channels: molecules ofmechanotransduction. J Cell Sci 117:2449–2460

99. Martinac B (2007) 3.5 billion years of mechanosensorytransduction: structure and function of mechanosensitive chan-nels in prokaryotes. In: Hamill OP (ed) Mechanosensitivechannels. Elsevier, Amsterdam (in press)

100. Martinac B, Buechner M, Delcour AH, Adler J, Kung C (1987)Pressure-sensitive ion channel in Escherichia coli. Proc NatlAcad Sci USA 84:2297–2301

101. Martinac B, Adler J, Kung C (1990) Mechanosensitive channelsof E. coli activated by amphipaths. Nature 348:261–263

102. Martinac B, Hamill OP (2002) Gramicidin A channels switchbetween stretch activation and stretch inactivation dependingupon bilayer thickness Proc Natl Acad Sci USA 99:4308–4312

103. Martinac B, Kloda A (2003) Evolutionary origins of mechano-sensitive ion channels. Prog Biophys Mol Biol 82:11–24

104. Mayer C, Maaser K, Daryab N, Zanker KS Broker EB, Friedl P(2004) Release of cell fragments by invading melanoma cells.Eur J Cell Biol 83:709–715

105. McBride DW Jr, Hamill OP (1999) A simplified fast pressure-clamp technique for studying mechanically-gated channels.Methods Enzymol 294:482–489

106. McBride DW Jr, Hamill OP (1995) A fast pressure clamptechnique for studying mechano-gated channels. In: Sakmann B,Neher E (eds) Single channel recording, 2nd edn. Plenum, NewYork, pp 329–340

107. McBride DW Jr, Hamill OP (1992) Pressure-clamp: a method forrapid step perturbation of mechanosensitive channels. PflugersArch 421:606–612

108. McBride DW Jr, Hamill OP (1993) Pressure-clamp technique formeasurement of the relaxation kinetics of mechanosensitivechannels. Trends Neurosci 16:341–345

109. McNeil PL, Steinhardt RA (2003) Plasma membrane disruption:repair, prevention, adaptation. Annu Rev Cell Dev Biol 19:697–731

110. McNeil PL, Teraski M (2001) Coping with the inevitable:how cells repair a torn surface membrane. Nat Cell Biol 3:E124–E129

111. McNeil PL, Kirchhausen T (2005) An emergency response teamfor membrane repair. Nat Rev Mol Cell Biol 6:499–505

112. Milton RL, Caldwell JH (1994) Membrane blebbing and tightseal formation: are there hidden artifacts in single channel patchclamp recordings? Comments Theor Biol 3:265–284

113. Moe P, Blount P (2005) Assessment of potential stimuli formechano-dependent gating of MscL: effects of pressure, tensionand lipid headgroups. Biochemistry 44:12239–12244

114. Morris CE (1990) Mechanosensitive ion channels. J Membr Biol113:93–107

115. Morris CE (2007) Mechanosensitivity of voltage-gated channels.In: Hamill OP (ed) Mechanosensitive channels. Elsevier,Amsterdam (in press)

116. Morris CE, Homann U (2001) Cell surface area regulation andmembrane tension. J Membr Biol 179:79–102

117. Morris CE, Horn R (1991) Failure to elicit neuronal macroscopicmechanosensitive currents anticipated by single-channel studies.Science 251:1246–1249

118. Morris CE, Sigurdson WJ (1989) Stretch-inactivated ion chan-nels coexist with stretch-activated channels. Science 243:807–809

119. Nichol JA, Hutter OF (1996) Tensile strength and dilatationalelasticity of giant sarcolemmal vesicles shed from rabbit muscle.J Physiol 493:187–198

120. O’Hagan R, Chalfie M, Goodman MB (2004) The MEC-4 DEG/ENaC channel of Caenorhabditis elegans touch receptor neuronstransduces mechanical signals. Nat Neurosci 8:43–50

121. Opsahl LR, Webb WW (1994) Lipid-glass adhesion in giga-sealed patch clamped membranes. Biophys J 66:75–79

122. Opsahl LR, Webb WW (1994) Transduction of membranetension by the ion channel alamethicin. Biophys J 66:71–74

123. Ostrow KL, Sachs F (2005) Mechanosensation and endothelin inastrocytes—hypothetical roles in CNS pathophysiology. BrainRes Brain Res Rev 48:488–508

124. Paoletti P, Ascher P (1994) Mechanosensitivity of NMDAreceptors in cultured mouse central neurons. Neuron 13:645–655

125. Patel AJ, Honoré E (2001) Properties and modulation ofmammalian 2P domain K+ channels. Trends Neurosci 24:339–346

126. Patel AJ, Honoré E, Maingret, F, Lesage F, Fink M, Duprat F,Lazdunski M (1998) A mammalian two pore domain mechano-gated S-like K+ channel. EMBO J 17:4283–4290

127. Patel AJ, Honoré E, Lesage F, Fink M, Romney G, Lazdunski M(1999) Inhalational anesthetics activated two-pore-domain back-ground K+ channels. Nat Neurosci 2:422–426

128. Patel AJ, Lazdunski M (2004) The 2P-domain K+ channels: rolein apoptosis and tumorigenesis. Pflugers Archiv 448:261–273

129. Patel AJ, Lazdunski M, Honoré E (2001) Lipid and mechano-gated 2P domain K+ channels. Curr Opin Cell Biol 13:422–428

130. Palmer LG (1984) Voltage-dependent block by amiloride andother monovalent cations of apical Na channels in the toadurinary bladder. J Membr Biol 80:153–165

131. Perozo E (2006) Gating prokaryotic mechanosensitive channels.Nat Rev Mol Cell Biol 7:109–119

132. Perozo E, Kloda A, Cortes DM, Martinac B (2002) Physicalprinciples underlying the transduction of bilayer deformationforces during mechanosensitive channel gating. Nat Struct Biol9:696–703

133. Perozo E, Rees DC (2003) Structure and mechanism inprokaryotic mechanosensitive channels. Curr Opin Struct Biol13:432–442

134. Prager GW, Binder BR (2004) Genetic determinants: is there anatherosclerosis gene. Acta Med Austriaca 31:1–7

135. Qi Z, Chi S, Su X, Naruse K, Sokabe M (2005) Activation of amechanosensitive BK channel by membrane stress created withamphipaths. Mol Membr Biol 22:519–527

136. Ramsey IS, Delling M, Clapham DE (2006) An introduction toTRP channels. Annu Rev Physiol 68:619–647

137. Rao JN et al (2006) TRPC1 functions as a store-operated Ca2+

channels in intestinal epithelial cells and regulates early mucosalrestitution after wounding. Am J Physiol 290:G782–G792

138. Raucher D, Sheetz MP (1999) Characteristic of a membranereservoir buffering membrane tension. Biophys J 77:1992–2002

139. Rees DC, Chang G, Spencer RH (2000) Crystallographicanalyses of ion channels: lessons and challenges. J Biol Chem275:713–716

140. Reifarth FW, Clauss W, Weber WM (1999) Stretch-independentactivation of the mechanosensitive cation channel in oocytes ofXenopus laevis. Biochim Biophys Acta 1417:63–76

141. Rüsch A, Kros CJ, Richardson GP (1994) Block by amilorideand its derivatives of mechano-electrical transduction in outerhair cells of mouse cochlear cultures. J Physiol 474:75–86

142. Ruknudin A, Song MJ, Sachs F (1991) The ultrastructure ofpatch-clamped membranes: a study using high voltage electronmicroscopy. J Cell Biol 112:125–134

143. Sachs F (1988) Mechanical transduction in biological systems.Crit Rev Biomed Eng 16:141–169

144. Sachs F, Morris CE (1988) Mechanosensitive ion channelsin nonspecialized cells. Rev Physiol Biochem Pharmacol 132:1–77

350 Pflugers Arch - Eur J Physiol (2006) 453:333–351

Page 19: Twenty odd years of stretch-sensitive channels · or osmotic pressure gradients applied across the patch without disrupting the seal. This feature allowed the first recordings of

145. Sachs F, Qin F (1993) Gated, ion selective channels observedwith patch pipettes in the absence of membranes: novelproperties of the giga-seal. Biophys J 65:1101–1107

146. Sackin H (1989) A stretch-activated K+ channel sensitive to cellvolume. Proc Natl Acad Sci USA 86:1731–1735

147. Sackin H (1995) Mechanosensitive channels. Annu Rev Physiol57:333–353

148. Sheetz MP, Dai J (1996) Modulation of membrane dynamics andcell motility by membrane tension. Trends Cell Biol 6:85–89

149. Sheetz MP, Sable JE, Döbereiner HG (2006) Continuousmembrane-cytoskeleton adhesion requires accommodation tolipid and cytoskeleton dynamics. Annu Rev Biophys BiomolStruct 35:417–434

150. Sigurdson WJ, Morris CE (1989) Stretch-activated ion channelin growth cones of snail neurons. J Neurosci 9:2801–2808

151. Small DL, Morris CE (1994) Delayed activation of singlemechanosensitive channels in Lymnaea neurons. Am J Physiol267:C598–C606

152. Sokabe M, Sachs F (1990) The structure and dynamics of patchclamped membrane: a study using differential interferencecontrast microscopy. J Cell Biol 111:599–606

153. Sokabe M, Sachs F, Jing Z (1991) Quantitative video microsco-py of patch clamped membranes—stress, strain, capacitance andstretch channel activation. Biophys J 59:722–728

154. Sokabe M, Nunogaki K, Naruse K, Soga H (1993) Mechanics ofpatch clamped and intact cell-membranes in relation to SAchannel activation (1993) Jpn J Physiol 43:S197–S204

155. Suchyna TM, Johnson JH, Hamer K, Leykam JF, Hage DA,Clemo HF, Baumgarten CM, Sachs F (1998) Identification of apeptide toxin from Grammostola spatula spider venom thatblocks cation selective stretch-activated channels. J Gen Physiol115:583–598

156. Suchyna TM, Besch SR, Sachs F (2004) Dynamic regulation ofmechanosensitive channels: capacitance used to monitor patchtension in real time. Phys Biol 1:1–18

157. Suchyna TM, Tape SE, Koeppe RE III, Anderson OS, Sachs F,Gottlieb PA (2004) Bilayer-dependent inhibition of mechano-sensitive channels by neuroactive peptide enatiomers. Nature430:235–240

158. Sukharev S (2002) Purification of the small mechanosensitivechannel in Escherichia coli (MscS): the subunit structure,conduction and gating characteristics. Biophys J 83:290–298

159. Sukharev SI, Blount P, Martinac B, Blattner FR, Kung C (1994)A large-conductance mechanosensitive channel in E. coliencoded by MscL alone. Nature 368:265–268

160. Sukahrev SI, Martinac B, Arshavsky VY, Kung C (1993) Two typesof mechanosensitive channels in the E. coli cell envelope:solubilization and functional reconstitution. Biophys J 65:177–183

161. Sukharev SI, Sigurdson WJ, Kung C, Sachs F (1999) Energeticand spatial parameters for gating of the bacterial large conduc-tance mechanosensitive channel, MscL. J Gen Physiol 113:525–539

162. Tanaka T, Tamba Y, Masum SM, Yamashita Y, Yamazaki M (2002)La3+ and Gd3+ induced shape change of giant unilamellar vesiclesof phosphatidylcholine. Biochim Biophys Acta 1564:173–182

163. Tang QY, Qi Z, Naruse K, Sokabe M (2003) Characterization ofa functionally expressed stretch-activated BKCa channel clonedfrom chick ventricular myocytes. J Membr Biol 196:185–200

164. Tongo T, Krasieva TB, Steinhardt RA (2000) A decrease inmembrane tension precedes successful cell-membrane repair.Mol Biol Cell 11:4339–4346

165. Tousova K, Vyklicky L, Susankova K, Benedikt J, Vlachova V(2004) Gadolinium activates and sensitizes the vanilloid receptor

TRPV1 through the external protonation sites. Mol Cell Neurosci30:207–217

166. Trebak M, Bird GS, Mckay RR, Putney JW Jr (2002)Comparison of human TRPC3 channels in receptor-activatedand store-operated modes. Differential sensitivity to channelblockers suggests fundamental differences in channel composi-tion. J Biol Chem 277:21617–21623

167. Vandebrouck C, Martin D, Colson-Van Schoor M, Debaix H,Gaily P (2002) Involvement of TRPC in the abnormal calciuminflux observed in dystrophic (mdx) mouse skeletal musclefibers. J Cell Biol 158:1089–1096

168. Vandorpe DH, Morris CE (1992) Stretch activation of theAplysia S-channel. J Membr Biol 127:205–214

169. Vanoye CG, Reuss L (1999) Stretch-activated single K+ channelsaccount for whole-cell currents elicited by swelling. Proc NatlAcad Sci USA 96:6511–6516

170. Verkhovsky AB, Svitkina TM, Borisy GG (1999) Self-polariza-tion and directional motility of cytoplasm. Curr Biol 9:11–20

171. Voets T, Talavera K, Owsiannik G, Nilius B (2005) Sensing withTRP channels. Nat Chem Biol 1:85–92

172. Vogel V, Sheetz M (2006) Local force and geometry sensingregulate cell functions. Nat Rev Mol Cell Biol 7:265–275

173. Wan X, Juranka P, Morris CE (1999) Activation of mechano-sensitive currents in traumatized membrane. Am J Physiol 276:C318–C327

174. Weber WM, Popp C, Clauss W, van Driessche W (2000)Maitotoxin induces insertion of different ion channels into theXenopus oocyte plasma membrane via Ca2+-stimulated exocyto-sis. Pflugers Archiv 439:363–369

175. Welsh DG, Morielli AD, Nelson MT, Brayden JE (2002)Transient receptor potential channels regulate myogenic tone ofresistance arteries. Circ Res 90:248–250

176. Wilkinson NS, Gao F, Hamill OP (1998) The effects of blockersof mechanically-gated channels on Xenopus oocyte growth anddevelopment. J Membr Biol 165:161–174

177. Wu G, McBride DW Jr, Hamill OP (1998) Mg2+ block andinward rectification of mechanosensitive channels in Xenopusoocytes. Pflugers Archiv 435:572–574

178. Yang XC, Sachs F (1989) Block of stretch-activated ion channelsin Xenopus oocytes by gadolinium and calcium ions. Science243:1068–1071

179. Yang XC, Sachs F (1990) Characterization of stretch-activatedion channels in Xenopus oocytes. J Physiol 431:103–122

180. Yeung EW, Allen DG (2004) Stretch-activated channels instretch-induced muscle damage: role in muscular dystrophy.Clin Exp Pharmacol Physiol 31:551–556

181. Zampighi GA, Kreman M, Boorer KJ, Loo DDF, Bezanilla F,Chandy G, Hall JE, Wright EM (1995) A method fordetermining the unitary functional capacity of cloned channelsand transporters expressed in Xenopus laevis oocytes. J MembrBiol 148:65–78

182. Zhang Y, McBride DW Jr, Hamill OP (1998) Ion selectivity of amembrane conductance activated by removal of extracellularcalcium in Xenopus oocytes. J Physiol 508:763–776

183. Zhang Y, Hamill OP (2000) Calcium, voltage and osmotic stresscurrents in Xenopus oocytes and their relationship to singlemechanically-gated channels. J Physiol 523.1:83–99

184. Zhang Y, Hamill OP (2000) On the discrepancy betweenmembrane patch and whole cell mechanosensitivity in Xenopusoocytes. J Physiol 523.1:101–115

185. Zhang Y, Gao F, Popov V, Wan J, Hamill OP (2000) Mechanically-gated channel activity in cytoskeleton deficient blebs and vesiclesfrom Xenopus oocytes. J Physiol 523.1:117–129

Pflugers Arch - Eur J Physiol (2006) 453:333–351 351